Skip to main content
Log in

Large Deviations, Dynamics and Phase Transitions in Large Stochastic and Disordered Neural Networks

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

Neuronal networks are characterized by highly heterogeneous connectivity, and this disorder was recently related experimentally to qualitative properties of the network. The motivation of this paper is to mathematically analyze the role of these disordered connectivities on the large-scale properties of neuronal networks. To this end, we analyze here large-scale limit behaviors of neural networks including, for biological relevance, multiple populations, random connectivities and interaction delays. Due to the randomness of the connectivity, usual mean-field methods (e.g. coupling) cannot be applied, but, similarly to studies developed for spin glasses, we will show that the sequences of empirical measures satisfy a large deviation principle, and converge towards a self-consistent non-Markovian process. From a mathematical viewpoint, the proof differs from previous works in that we are working in infinite-dimensional spaces (interaction delays) and consider multiple cell types. The limit obtained formally characterizes the macroscopic behavior of the network. We propose a dynamical systems approach in order to address the qualitative nature of the solutions of these very complex equations, and apply this methodology to three instances in order to show how non-centered coefficients, interaction delays and multiple populations networks are affected by disorder levels. We identify a number of phase transitions in such systems upon changes in delays, connectivity patterns and dispersion, and particularly focus on the emergence of non-equilibrium states involving synchronized oscillations.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Coupling methods, introduced by Dobrushin [15] and very well documented in the book of Sznitman [36], consist in exhibiting a particular process, solution of the limit equations which is the almost sure limit of the network equations. This process is constructed explicitly by defining the solution of the mean-field equations with identical Brownian motions and initial conditions as used in the network equations.

  2. We only need to assume that S takes values in a bounded interval of \(\mathbb{R}\), which without loss of generality (by scaling and translation of the variables x j) can be mapped to [0,1].

  3. The solution to (1) therefore lives on the product space Ω×Ω′.

  4. Remark that these processes only depend on the parameters (mean and standard deviation) of the weights J ij and on the law of the solution, and not on the realization of the weights J ij . Hence this formulation is not inconsistent with Boltzmann’s stoßzahlansatz.

  5. This can appear as a strong assumption. However, recent experimental researches tend to show that such states where neurons are independent are naturally present in the brain [16, 31].

  6. This property simply expresses that the dispersion of the solution is an increasing function of the disorder parameters. Numerical simulations in different situations are provided in Fig. 3.

  7. For instance, during the revision of this article, a preprint appeared [17] where the authors used similar techniques as ours to treat a case with slightly different synaptic models in a single-populations network with discrete-time dynamics and no delay.

References

  1. Amari, S.: Characteristics of random nets of analog neuron-like elements. IEEE Trans. Syst. Man Cybern. 2(5), 643–657 (1972). 1972

    Article  MathSciNet  MATH  Google Scholar 

  2. Aradi, I., Soltesz, I.: Modulation of network behaviour by changes in variance in interneuronal properties. J. Physiol. 538, 227 (2002)

    Article  Google Scholar 

  3. Ben-Arous, G., Guionnet, A.: Large deviations for Langevin spin glass dynamics. Probab. Theory Relat. Fields 102, 455–509 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bressloff, P.C.: Spatiotemporal dynamics of continuum neural fields. J. Phys. A, Math. Theor. 45, 033001 (2012)

    Article  MathSciNet  ADS  Google Scholar 

  5. Buice, M., Cowan, J., Chow, C.: Systematic fluctuation expansion for neural network activity equations. Neural Comput. 22, 377–426 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  6. Burton, R.M., Dehling, H.: Large deviations for some weakly dependent random processes. Stat. Probab. Lett. 9, 397–401 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  7. Buzsaki, G.: Rhythms of the Brain. Oxford University Press, London (2004)

    Google Scholar 

  8. Candes, E.J., Tao, T.: Near-optimal signal recovery from random projections: universal encoding strategies? IEEE Trans. Inf. Theory 52, 5406–5425 (2006)

    Article  MathSciNet  Google Scholar 

  9. Cessac, B., Doyon, B., Quoy, M., Samuelides, M.: Mean-field equations, bifurcation map and route to chaos in discrete time neural networks. Phys. D, Nonlinear Phenom. 74, 24–44 (1994)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  10. Crisanti, A., Sompolinsky, H.: Dynamics of spin systems with randomly asymmetric bounds: Ising spins and Glauber dynamics. Phys. Rev. A 37, 4865 (1987)

    Article  MathSciNet  ADS  Google Scholar 

  11. Da Prato, G., Zabczyk, J.: Stochastic Equations in Infinite Dimensions. Cambridge University Press, Cambridge (1992)

    Book  MATH  Google Scholar 

  12. Dauce, E., Moynot, O., Pinaud, O., Samuelides, M.: Mean-field theory and synchronization in random recurrent neural networks. Neural Process. Lett. 14, 115–126 (2001)

    Article  MATH  Google Scholar 

  13. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications, vol. 38. Springer, Berlin (2009)

    Google Scholar 

  14. Deuschel, J., Stroock, D.: Large Deviations, vol. 137. Academic Press, San Diego (1989)

    MATH  Google Scholar 

  15. Dobrushin, R.: Prescribing a system of random variables by conditional distributions. Theory Probab. Appl. 15 (1970)

  16. Ecker, A., Berens, P., Keliris, G., Bethge, M., Logothetis, N., Tolias, A.: Decorrelated neuronal firing in cortical microcircuits. Science 327, 584 (2010)

    Article  ADS  Google Scholar 

  17. Faugeras, O., MacLaurin, J.: A large deviation principle for networks of rate neurons with correlated synaptic weights (2013). Preprint, arXiv:1302.1029

  18. Faugeras, O., Touboul, J., Cessac, B.: A constructive mean-field analysis of multi population neural networks with random synaptic weights. In: COSYNE 09 (2009)

    Google Scholar 

  19. Guionnet, A.: Averaged and quenched propagation of chaos for spin glass dynamics. Probab. Theory Relat. Fields 109, 183–215 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  20. Hale, J., Lunel, S.: Introduction to Functional Differential Equations. Springer, Berlin (1993)

    Book  MATH  Google Scholar 

  21. Hopfield, J.J.: Neural networks and physical systems with emergent collective computational abilities. Proc. Natl. Acad. Sci. USA 79, 2554–2558 (1982)

    Article  MathSciNet  ADS  Google Scholar 

  22. Jaeger, H.: Echo state network. Scholarpedia 2, 2330 (2007)

    Article  ADS  Google Scholar 

  23. Kandel, E., Schwartz, J., Jessel, T.: Principles of Neural Science, 4th edn. McGraw-Hill, New York (2000)

    Google Scholar 

  24. Kaufman, M., Corner, M.A., Ziv, N.E.: Long-term relationships between cholinergic tone, synchronous bursting and synaptic remodeling. PLoS ONE 7, e40980 (2012)

    Article  ADS  Google Scholar 

  25. Mao, X.: Stochastic Differential Equations and Applications. Horwood Publishing, Chichester (2008)

    Book  Google Scholar 

  26. Marder, E., Goaillard, J.: Variability, compensation and homeostasis in neuron and network function. Nat. Rev. Neurosci. 7, 563–574 (2006)

    Article  Google Scholar 

  27. Moynot, O.: Etude mathématique de la dynamique des réseaux neuronaux aléatoires récurrents. PhD thesis (2000)

  28. Moynot, O., Samuelides, M.: Large deviations and mean-field theory for asymmetric random recurrent neural networks. Probab. Theory Relat. Fields 123, 41–75 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  29. Parker, D.: Variable properties in a single class of excitatory spinal synapse. J. Neurosci. 23, 3154–3163 (2003)

    Google Scholar 

  30. Rajan, K., Abbott, L.F., Sompolinsky, H.: Stimulus-dependent suppression of chaos in recurrent neural networks. Phys. Rev. E 82, 011903 (2010)

    Article  ADS  Google Scholar 

  31. Renart, A., De la Rocha, J., Bartho, P., Hollender, L., Parga, N., Reyes, A., Harris, K.: The asynchronous state in cortical circuits. Science 327, 587 (2010)

    Article  ADS  Google Scholar 

  32. Roxin, A., Brunel, N., Hansel, D.: Role of delays in shaping spatiotemporal dynamics of neuronal activity in large networks. Phys. Rev. Lett. 94, 238103 (2005)

    Article  ADS  Google Scholar 

  33. Sompolinsky, H., Crisanti, A., Sommers, H.: Chaos in random neural networks. Phys. Rev. Lett. 61, 259–262 (1988)

    Article  MathSciNet  ADS  Google Scholar 

  34. Sompolinsky, H., Zippelius, A.: Relaxational dynamics of the Edwards-Anderson model and the mean-field theory of spin-glasses. Phys. Rev. B 25, 6860–6875 (1982)

    Article  ADS  Google Scholar 

  35. Sznitman, A.: Equations de type de Boltzmann, spatialement homogenes. Probab. Theory Relat. Fields 66, 559–592 (1984)

    MathSciNet  MATH  Google Scholar 

  36. Sznitman, A.: Topics in propagation of chaos. In: Ecole d’Eté de Probabilités de Saint-Flour XIX – 1989. Lecture Notes in Mathematics, vol. 1464, pp. 165–251 (1991)

    Chapter  Google Scholar 

  37. Touboul, J.: Mean-field equations for stochastic firing-rate neural fields with delays: derivation and noise-induced transitions. Phys. D, Nonlinear Phenom. 241(15), 1223–1244 (2012)

    Article  MathSciNet  ADS  MATH  Google Scholar 

  38. Touboul, J.: Propagation of chaos in neural fields. Ann. Appl. Probab. (2013) (in press)

  39. Touboul, J., Hermann, G.: Heterogeneous connections induce oscillations in large-scale networks. Phys. Rev. Lett. (2012) (in press). doi:10.1103/PhysRevLett.109.018702

    Google Scholar 

  40. Touboul, J., Hermann, G., Faugeras, O.: Noise-induced behaviors in neural mean field dynamics. SIAM J. Dyn. Syst. 11 (2011). doi:10.1137/110832392

  41. Toyoizumi, T., Abbott, L.: Beyond the edge of chaos: amplification and temporal integration by recurrent networks in the chaotic regime. Phys. Rev. E 84, 051908 (2011)

    Article  ADS  Google Scholar 

  42. Triller, A., Choquet, D.: New concepts in synaptic biology derived from single-molecule imaging. Neuron 59, 359–374 (2008)

    Article  Google Scholar 

  43. Wainrib, G., Touboul, J.: Topological and dynamical complexity of random neural networks. Phys. Rev. Lett. (2012) (in press). doi:10.1103/PhysRevLett.110.118101

    Google Scholar 

  44. Wilson, C.: Up and down states. Scholarpedia 3, 1410 (2008)

    Article  ADS  Google Scholar 

  45. Wilson, H., Cowan, J.: Excitatory and inhibitory interactions in localized populations of model neurons. Biophys. J. 12, 1–24 (1972)

    Article  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jonathan Touboul.

Additional information

INRIA BANG Laboratory and the Mathematical Neuroscience Lab, CIRB-Collège de France, CNRS UMR 7241 INSERM U1050, Université Pierre et Marie Curie ED 158. MEMOLIFE Laboratory of excellence and Paris Sciences Lettres PSL*.

Appendices

Appendix A: H is a Good Rate Function

This appendix is devoted to the proof of the Proposition 3 stating that the map H given by (20) is a good rate function, namely that it is lower semi-continuous with compact level sets. In order to perform this demonstration, it is convenient to analyze for a moment an intermediate system where the interaction is discretized in time, which will expedite the analysis of our continuous time problem.

Given an integer k, we define Δk={0=t 0<t 1<⋯<t k <t k+1=T} a partition of [0,T] and consider the following dynamics for the neurons of population α:

$$\begin{aligned} \begin{cases} dx^i_t= (-\frac{1}{\theta_{\alpha}}x^i_t + \sum_{\gamma=1}^M \sum_{j: p(j)=\gamma} J_{ij} S_{\alpha p(j)}(x^j_{t^{(k)}-\tau_{\alpha p(j)}}) )\,dt+ \lambda_{\alpha}dW^{i}_t\\ t^{(k)}= \sup\{ t_l \in\Delta^k | t_l \leq t \}\\ \mbox{Law of } (x_t^{\alpha})_{t\in[-\tau,0]} = \mu_{\alpha }^{\otimes N_{\alpha}}. \end{cases} \end{aligned}$$
(33)

As in Proposition 1, this system clearly admits a unique weak solution for any \(J \in\mathbb{R}^{N \times N}\). We will denote Q n,α,(k)(J) its restriction to the σ-algebra \(\sigma(x^{i}_{s}, 1\leq i \leq N ; p(i)=\alpha, -\tau\leq s\leq T)\), and \(Q^{n,\alpha,(k)}=\mathcal{E}(Q^{n,\alpha,(k)}(J))\). They are both probability measures on \(\mathcal{C}([-\tau,T],\mathbb{R})^{n}\). By Girsanov Theorem, \(Q^{n,\alpha,(k)}(J) \ll P_{\alpha}^{\otimes n}\) with:

$$\begin{aligned} \frac {\mbox {d} Q^{n,\alpha,(k)}(J)}{\mbox {d} {P_{\alpha}}^{\otimes n}} =& \exp\Biggl\{ \sum_{i:p(i)=\alpha} \int_{0}^{T} \Biggl(\frac{1}{\lambda_{\alpha}} \sum_{\gamma=1}^{P} \sum_{j:p(j)=\gamma} J_{ij} S_{\alpha\gamma} \bigl(x_{t^{(k)}-\tau _{\alpha\gamma}}^j\bigr) \Biggr) dW_{t}^{i} \\ &{}- \int_{0}^{T} \Biggl(\frac{1}{\lambda_{\alpha}} \sum _{\gamma =1}^{P} \sum _{j:p(j)=\gamma} J_{ij} S_{\alpha\gamma}\bigl(x_{t^{(k)}-\tau_{\alpha \gamma}}^j \bigr) \Biggr)^2 dt \Biggr\}. \end{aligned}$$

We define for \(\mu\in\mathcal{M}_{1}^{+}(\mathcal{C})\) the discrete-time version of Γ(μ), denoted Γ k(μ) on Δk, as:

$$\begin{aligned} \varGamma^{k}(\mu) =& \int_{\mathcal{C}} \log\Biggl( \int \exp\Biggl\{ \sum_{l=0}^k \bigl( \mathbf{G}_{t_l}(\omega)+\mathbf{m}_{\mu}(t_l) \bigr)' \cdot(\mathbf{W} _{t_{l+1}} - \mathbf{W} _{t_l} ) (x) \\ &{}- \frac{1}{2} \sum_{l=0}^k \bigl\| \mathbf{G}_{t_l}(\omega)+\mathbf{m} _{\mu}(t_l) \bigr\|^2(t_{l+1}-t_l) \Biggr\} d \gamma_{K_{\mu}}(\omega) \Biggr) d\mu(x). \end{aligned}$$

Since the covariance K μ is diagonal and therefore the components of G under γ μ are independent, this map is the sum of the functions Γ α,k given by:

$$\begin{aligned} \varGamma^{\alpha,k}(\mu) =& \int_{\mathcal{C}} \log\Biggl( \int \exp\Biggl\{ \sum_{l=0}^k \bigl(G_{t_l}^{\alpha}(\omega)+m_{\mu}^{\alpha}(t_l) \bigr) \bigl(W^{\alpha}_{t_{k+1}} - W^{\alpha}_{t_l} \bigr) (x) \\ &{}- \frac{1}{2} \sum_{l=0}^k \bigl(G_{t_l}^{\alpha}(\omega)+m_{\mu}^{\alpha}(t_l) \bigr)^2(\omega) (t_{l+1}-t_l) \Biggr\} d \gamma_{K_{\mu}}(\omega) \Biggr) d\mu(x). \end{aligned}$$

The discretized version of our function H of interest is simply:

$$\begin{aligned} H^k (\mu) = \left\{ \begin{array}{l@{\quad}l} I(\mu|P) - \varGamma^k(\mu) & \mbox{if } I(\mu|P) < \infty,\\ \infty& \mbox{otherwise}. \end{array} \right. \end{aligned}$$

We will show that the map H k is a good rate function. This proof proceeds by introducing additional maps on \(\mathcal{M}_{1}^{+}(\mathcal{C})\):

$$\begin{aligned} \varGamma_1^k(\mu) =& \log\biggl( \int\exp\biggl( - \frac{1}{2} \int_0^T \bigl\| \mathbf{G}_{t^{(k)}}(\omega) \bigr\|^2dt \biggr) d\gamma_{K_{\mu }}(\omega) \biggr) - \frac{1}{2} \int _0^T \bigl\|\mathbf{m}_{\mu} \bigl(t^{(k)}\bigr) \bigr\|^2 dt\\ \varGamma_2^k(\mu) =& \frac{1}{2} \int\int\biggl( \int_0^T \mathbf{G}_{t^{(k)} }' \cdot d\mathbf{W}_t(x) - \mathbf{m}_{\mu} \bigl(t^{(k)}\bigr)dt\biggr) ^2 d\gamma_{\widetilde {K}_{\mu}^{T,k}} d \mu(x) \\ &{}+\int\int\mathbf{m}_{\mu}\bigl(t^{(k)}\bigr)' \cdot d\mathbf{W}_t (x) d\mu(x) \end{aligned}$$

where

$$\begin{aligned} &\widetilde{K}_{\mu}^{t,k}(s,u)\\ &\quad{}= \biggl(\int\frac{ \exp\{ - \frac {1}{2} \int_0^t (G^{\alpha}_{u^{(k)}}(\omega) )^2+\mathbf {1}_{\alpha \neq\gamma} (G^{\gamma}_{u^{(k)}}(\omega) )^2du \} G^{\gamma}_{u^{(k)} }(\omega)G^{\alpha}_{s^{(k)}}(\omega) }{\int\exp\{ - \frac {1}{2} \int_0^t (G^{\alpha}_{u^{(k)}}(\omega) )^2+\mathbf{1}_{\alpha\neq \gamma } (G^{\gamma}_{u^{(k)}}(\omega) )^2du \} d\gamma_{\mu}} d\gamma_{\mu} \biggr)_{\alpha, \gamma\in\{1 \cdots M\}}. \end{aligned}$$

One can easily see that this function takes values in the M×M diagonal positive matrices. Moreover, we can define \(\gamma _{\widetilde {K}_{\mu}^{T,k}}\), probability measure on Ω, such that

$$\begin{aligned} d\gamma_{\widetilde{K}_{\mu}^{T,k}}= \frac{\prod_{\alpha=1}^M \exp \{ - \frac{1}{2} \int_0^T (G^{\alpha}_{t^{(k)}}(\omega) )^2 dt \} }{\int\prod_{\alpha=1}^M \exp\{ - \frac{1}{2} \int_0^T (G^{\alpha}_{t^{(k)}}(\omega) )^2 dt \} d\gamma_{\mu}} d\gamma_{\mu}, \end{aligned}$$

under which G is a M-dimensional centered Gaussian process with covariance \(\widetilde{K}_{\mu}^{T,k}\) (this Gaussian calculus property is proved for instance in [3, Appendix A]).

Proposition 4

$$\begin{aligned} \varGamma^{k}(\mu)= \varGamma_1^{k}(\mu) + \varGamma_2^{k}(\mu) \end{aligned}$$

Proof

Let

$$\begin{aligned} \varGamma_1^{\alpha,k}(\mu) =& \log\biggl( \int\exp\biggl( - \frac {1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2( \omega)dt \biggr) d\gamma_{K_{\mu }}(\omega) \biggr) - \frac{1}{2} \int _0^T \bigl(m_{\mu}^{\alpha} \bigl(t^{(k)}\bigr)\bigr)^2 dt,\\ \varGamma_2^{\alpha,k}(\mu) =& \frac{1}{2} \int\int\biggl( \int_0^T G_{t^{(k)} }^{\alpha} \bigl(dW_t^{\alpha}(x) - m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\gamma _{\widetilde{K}_{\mu}^{T,k}} d\mu(x)\\ &{} + \int\int m_{\mu}^{\alpha }\bigl(t^{(k)} \bigr)dW^{\alpha}_t (x) d\mu(x). \end{aligned}$$

For \(\varGamma_{i}^{k}= \sum_{\alpha=1}^{M} \varGamma_{i}^{\alpha,k}, i\in \{1,2\}\), it is sufficient to prove that

$$\begin{aligned} \varGamma^{\alpha,k}(\mu)= \varGamma_1^{\alpha,k}(\mu) + \varGamma_2^{\alpha ,k}(\mu). \end{aligned}$$

Simple manipulations yield

$$\begin{aligned} \varGamma^{\alpha, k}(\mu) = & \int\log\biggl( \int\exp\biggl\{ \int _0^T \bigl(G^{\alpha}_{t^{(k)}}( \omega)+m_{\mu}^{\alpha}\bigl(t^{(k)}\bigr) \bigr)dW_t^{\alpha }(x) \\ &{} - \frac{1}{2} \int_0^T \bigl(G_{t^{(k)}}^{\alpha}(\omega)+m_{\mu }^{\alpha } \bigl(t^{(k)}\bigr) \bigr)^2 dt \biggr\} d \gamma_{K_{\mu}}(\omega) \biggr) d\mu(x) \\ = & \int\log\biggl\{ \biggl( \exp\biggl\{-\frac{1}{2} \int _0^T \bigl(m^{\alpha }_{\mu} \bigl(t^{(k)}\bigr)\bigr)^2dt \biggr\} \int\exp\biggl\{- \frac{1}{2}\int_0^T \bigl(G^{\alpha }_{t^{(k)}} \bigr)^2dt \biggr\} d\gamma_{\mu} \biggr) \\ &{} \times\biggl( \exp\biggl\{ \int_0^T m_{\mu}^{\alpha}\bigl(t^{(k)}\bigr) dW_t^{\alpha } \biggr\} \\ &{} \times\int\exp\biggl\{ \int_0^T G_{t^{(k)}}^{\alpha} \bigl(dW_t^{\alpha} - m_{\mu}^{\alpha}\bigl(t^{(k)}\bigr)dt \bigr) \biggr\} d \gamma_{\widetilde {K}_{\mu}^{T,k}} \biggr) \biggr\} d\mu \\ = & \log\biggl\{ {\mathcal{E}}_{\mu} \biggl[\exp{ \biggl( - \frac {1}{2} \int_0^T \bigl(G^{\alpha}_{t^{(k)}}\bigr)^2 dt \biggr) } \biggr] \biggr\} - \frac {1}{2}\int_0^T \bigl(m^{\alpha}_{\mu}\bigl(t^{(k)}\bigr) \bigr)^2 dt \\ &{} + \int\int_0^T m_{\mu}^{\alpha}\bigl(t^{(k)}\bigr) dW_t^{\alpha} d\mu \\ &{} +\int\log\biggl\{ \int\exp{ \biggl(\int_0^T G_{t^{(k)}}^{\alpha } \bigl(dW_t^{\alpha} - m_{\mu}^{\alpha}\bigl(t^{(k)}\bigr)dt \bigr) \biggr) } d \gamma_{\widetilde{K}_{\mu}^{T,k}} \biggr\} d\mu, \end{aligned}$$

and standard Gaussian calculus implies that

$$\begin{aligned} &\int\exp{ \biggl(\int_0^T G_{t^{(k)}}^{\alpha} \bigl(dW_t^{\alpha} - m_{\mu }^{\alpha} \bigl(t^{(k)}\bigr)dt \bigr) \biggr) } d\gamma_{\widetilde{K}_{\mu }^{T,k}}\\ &\quad{} = \exp \biggl\{ \frac{1}{2} \int\biggl( \int_0^T G_{t^{(k)}}^{\alpha} \bigl(dW_t^{\alpha}-m_{\mu}^{\alpha} \bigl(t^{(k)}\bigr)dt \bigr) \biggr)^2 d\gamma _{\widetilde {K}_{\mu}^{T,k}} \biggr\}, \end{aligned}$$

so that

$$\begin{aligned} \varGamma^{\alpha, k}(\mu) = &\varGamma^{\alpha, k}_1(\mu) + \int\int_0^T m_{\mu}^{\alpha} \bigl(t^{(k)}\bigr) dW_t^{\alpha} d\mu \\ &{}+ \int\log \biggl\{ \exp\biggl\{ \frac{1}{2} \int\biggl( \int_0^T G_{t^{(k)} }^{\alpha} \bigl(dW_t^{\alpha}-m_{\mu}^{\alpha} \bigl(t^{(k)}\bigr)dt \bigr) \biggr)^2 d\gamma _{\widetilde{K}_{\mu}^{T,k}} \biggr\} \biggr\} d\mu \\ = &\varGamma_1^{\alpha, n}(\mu) + \int\int _0^T m_{\mu}^{\alpha } \bigl(t^{(k)}\bigr) dW_t^{\alpha} d\mu\\ &{}+ \frac{1}{2} \int\int\biggl( \int_0^T G_{t^{(k)} }^{\alpha} \bigl(dW_t^{\alpha}-m_{\mu}^{\alpha} \bigl(t^{(k)}\bigr)dt \bigr) \biggr)^2 d\gamma_{\widetilde{K}_{\mu}^{T,k}} d\mu \end{aligned}$$

which concludes the proof. □

In order to show that H k is a good rate function, we need to resort to the definition of two maps defined, for \(\mu, \nu\in\mathcal {M}_{1}^{+}(\mathcal{C} )\), by:

$$\begin{aligned} \varGamma^{k}_{\nu}(\mu) = \int_{\mathcal{C}} \log\biggl( \int\exp\biggl\{ \int_0^T \bigl( \mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\nu}\bigl(t^{(k)} \bigr) \bigr)' \cdot d\mathbf{W} _t(x) \\ - \frac{1}{2} \int_0^T \bigl\| \mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\nu } \bigl(t^{(k)}\bigr) \bigr\| ^2dt \biggr\} d\gamma_{K_{\nu}}( \omega) \biggr) d\mu(x) \end{aligned}$$

and

$$\begin{aligned} \varGamma_{2,\nu}^{\alpha, k}(\mu) =& \frac{1}{2} \int\int \biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu}\bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\gamma_{\widetilde{K}_{\nu}^{T,n}} d\mu+ \int\int m^{\alpha }_{\nu } \bigl(t^{(k)}\bigr)dW^{\alpha}_t d\mu, \end{aligned}$$

which allows to define \(\varGamma_{2,\nu}^{k}\) as the sum over α of \(\varGamma_{2,\nu}^{\alpha, k}\) and

$$\begin{aligned} \varGamma_{\nu}^{\alpha,k}=\varGamma_1^{\alpha,k}+ \varGamma_{2,\nu}^{\alpha,k}. \end{aligned}$$

One can easily see that \(\varGamma^{k}_{\nu} = \sum_{\alpha=1}^{M} \varGamma _{\nu}^{\alpha, k}\). We eventually introduce a modified H-function:

$$\begin{aligned} H^{k}_{\nu} : \mathcal{M}_1^+(\mathcal{C}) \rightarrow&\mathbb{R}^+ \\ \mu \mapsto&\begin{cases} I(\mu|P) - \varGamma_{\nu}^{k}(\mu) & \mbox{if } I(\mu|P) < \infty,\cr \infty& \mbox{otherwise}. \end{cases} \end{aligned}$$

All these functions enjoy the following properties:

Lemma 5

We recall that d T denotes the Vaserstein distance (21).

  1. 1.

    There exists a positive constant C T , depending on T but not on n, such that: \(|\varGamma_{1}^{k}(\mu)-\varGamma_{1}^{k}(\nu)| \leq C_{T} d_{T}(\mu,\nu)\).

  2. 2.

    Γ kI(.|P) i.e. H k is a positive function. In particular, Γ k is finite whenever I(.|P) is.

  3. 3.

    There exists real constants a<1 and η>0 such that Γ kaI(.|P)+η.

  4. 4.

    There exists a positive constant C T , depending on T but not on n, such that: \(|\varGamma_{2,\nu}^{k}(\mu)-\varGamma_{2}^{k}(\mu)| \leq C_{T} (1+I(\mu|P) ) d_{T}(\mu,\nu)\).

  5. 5.

    Defining the following probability measure on \(\mathcal {M}_{1}^{+}(\mathcal{C})\):

    $$\begin{aligned} dQ_{\nu}^k(x) = &\exp{\varGamma_{\nu}^{k}( \delta_x)}dP(x) \\ = &\int\exp \biggl( \int_0^T \bigl( \mathbf{G}_{t^{(k)}} + \mathbf{m}_{\nu}\bigl(t^{(k)} \bigr) \bigr)' \cdot\, d\mathbf{W}_t(x)\\ &{} - \frac{1}{2} \int_0^T \bigl\| \mathbf {G}_{t^{(k)}} + \mathbf{m}_{\nu}\bigl(t^{(k)} \bigr) \bigr\|^2 dt \biggr) \, d\gamma_{\nu} \, dP(x) \end{aligned}$$

    we have \(H_{\nu}^{k}=I(.|Q_{\nu}^{k})\), so that \(H_{\nu}^{k}\) is lower semi-continuous on \(\mathcal{M}_{1}^{+}(\mathcal{C})\).

  6. 6.

    H k is a good rate function.

Once this lemma is proved, it will be easy to demonstrate our Proposition 3 stating that H is a good rate function. In details, we will show that:

Proposition 5

As the partition Δk is refined, we have:

  1. 1.

    On the compact set \(K_{L} = \{\mu\in\mathcal{M}_{1}^{+}(\mathcal {C})| I(\mu|P) \leq L \} \), Γ k converges uniformly to Γ.

  2. 2.

    \(\forall\mu\in K_{L}, \; \varGamma^{\alpha}(\mu) = \varGamma ^{\alpha }_{1}(\mu) + \varGamma^{\alpha}_{2}(\mu)\) where

  3. 3.

    ΓI(.|P) anda<1,η>0| ΓaI(.|P)+η.

  4. 4.

    H is a good rate function.

Proof

The proof of Lemma 5 (that follows) is easily extended to the continuous case.

(i), (ii):

Similarly to the proof of the assertions (i) and (iv) of Lemma 5, we can find two constants C 1,T and C 2,T such that

$$ \begin{aligned} \bigl| \bigl(\varGamma_1^{\alpha, k}-\varGamma_1^{\alpha, k+p} \bigr) (\mu) \bigr| & \leq C_{1,T} \max_{\gamma=1,M} \biggl( \int\int_0^T \bigl|S_{\alpha \gamma} \bigl(x^{\gamma }_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\\ &\quad{} - S_{\alpha\gamma} \bigl(x^{\gamma }_{t^{(k+p)}-\tau_{\alpha\gamma}}\bigr) \bigr|^2 dt \, d\mu(x) \biggr)^{\frac{1}{2}} \\ & \leq C_{1,T} \sqrt{T} \max_{\gamma=1,M} \biggl( \int\sup _{|t-s| \leq |\Delta_k|} \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_t \bigr) - S_{\alpha\gamma }\bigl(x^{\gamma}_s\bigr) \bigr|^2 d\mu(x) \biggr)^{\frac{1}{2}} \\ \bigl| \bigl(\varGamma_2^{\alpha,k}-\varGamma_2^{\alpha, k+p} \bigr) (\mu) \bigr| & \leq C_{2,T} \bigl(I(\mu|P) +1 \bigr) \max _{\gamma=1,M} \biggl( \int\sup_{|t-s| \leq|\Delta_k|} \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_t\bigr)\\ &\quad{} - S_{\alpha\gamma} \bigl(x^{\gamma }_s\bigr) \bigr|^2 d\mu(x) \biggr)^{\frac{1}{2}}. \end{aligned} $$

But, according to (22), we have for any a≥0

$$\begin{aligned} &a \int\sup_{|t-s| \leq|\Delta_k|} \bigl|S_{\alpha\gamma }\bigl(x^{\gamma}_t \bigr) - S_{\alpha\gamma} \bigl(x^{\gamma}_s\bigr) \bigr|^2 d\mu(x)\\ &\quad{}\leq I(\mu|P) + \log\int\exp{ \Bigl\{ a \sup _{|t-s| \leq|\Delta_k|} \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_t \bigr) - S_{\alpha\gamma} \bigl(x^{\gamma }_s\bigr) \bigr|^2 \Bigr\}} dP(x). \end{aligned}$$

The bounded convergence theorem ensures that

$$\begin{aligned} \lim_{k\to\infty} \log{\int\exp{ \Bigl\{ a \sup_{|t-s| \leq |\Delta _k|} \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_t\bigr) - S_{\alpha\gamma } \bigl(x^{\gamma}_s\bigr) \bigr|^2 \Bigr\}} dP(x) } = 0. \end{aligned}$$

Let ε>0, choosing \(a=\frac{1}{\varepsilon^{2}}\), it is easy to see that there exists an integer k(ε) such that, for kk(ε),∀γ∈{1,…M}:

$$\begin{aligned} \int\sup_{|t-s| \leq|\Delta_k|} \bigl|S_{\alpha\gamma}\bigl(x^{\gamma }_t \bigr) - S_{\alpha\gamma} \bigl(x^{\gamma}_s\bigr) \bigr|^2 d\mu(x) \leq\bigl(I(\mu|P) + 1 \bigr) \varepsilon^2. \end{aligned}$$

Hence, for any kk(ε), any p, and any μK L :

$$\begin{aligned} \bigl| \bigl(\varGamma_1^{\alpha,k}-\varGamma_1^{\alpha,k+p} \bigr) (\mu) \bigr| \leq &C_T (1+L )^{\frac{1}{2}}\varepsilon \\ \bigl| \bigl(\varGamma_2^{\alpha,k}-\varGamma_2^{\alpha,k+p} \bigr) (\mu)\bigr | \leq& C_T (1+L )^{\frac{3}{2}}\varepsilon. \end{aligned}$$

Which shows that \(\varGamma_{1}^{\alpha,k}\), \(\varGamma_{2}^{\alpha ,k}\), and thus Γ α,k converge uniformly on K L . It is not difficult to see that the respective limits are \(\varGamma_{1}^{\alpha}\), \(\varGamma _{2}^{\alpha}\) and Γ α, which implies \(\varGamma ^{\alpha }=\varGamma_{1}^{\alpha}+\varGamma_{2}^{\alpha}\) on K L . Besides, as \(\varGamma^{k} = \sum_{\alpha=1}^{M} \varGamma^{\alpha,k}\) and \(\varGamma= \sum_{\alpha=1}^{M} \varGamma^{\alpha}\), we also have the uniform convergence of Γ k towards Γ on K L .

(iii):

This can be proved exactly as that the properties (iii) and (iv) of Lemma 5.

(iv):

We show that {HL} is a compact set. H≥(1−a)I(|P)−η so that \(I(|P) \leq\frac{H+\eta}{1-a}\). Hence \(\{H \leq L \}\subset\{I(|P) \leq\frac{L+\eta }{1-a} \}\). Let \((\mu_{p})_{p} \in\{H \leq L \}^{\mathbb{N}} \subset\{I(|P) \leq\frac{L+\eta}{1-a} \}^{\mathbb{N}}\). As here \(\{I(|P) \leq\frac{L+\eta}{1-a} \}\) is a compact set, there exists a subsequence \((\mu_{p_{m}})_{m}\) such that \(\mu_{p_{m}} \to \mu \) as m→∞. We conclude by stating that, as H k converge uniformly towards H on \(\{I(|P) \leq\frac{L+\eta}{1-a} \}\), the latest inherits the lower semi-continuity of the firsts. Hence {HL} is a closed set so that μ∈{HL} and \((\mu_{p_{m}})_{m}\) converges in {HL}.

 □

We are hence left proving Lemma 5.

Proof of Lemma 5.(i)

We prove Lipschitz-continuity for \(\varGamma^{\alpha, k}_{1}\). We have:

$$\begin{aligned} &\biggl|\log\biggl( 1 + \frac{\int\exp( - \frac{1}{2} \sum_{l=0}^k {G^{\alpha}_{t_l}}^2(t_{l+1} - t_l) ) d(\gamma_{K_{\mu}}-\gamma _{K_{\nu}})}{\int\exp( - \frac{1}{2} \sum_{l=0}^k {G^{\alpha }_{t_l}}^2(t_{l+1} - t_l) ) d\gamma_{K_{\nu}}} \biggr) \biggr| \\ &\quad{}= \biggl|\varGamma _1^{\alpha,k}( \mu)-\varGamma_1^{\alpha,k}(\nu)+\frac {1}{2}\int\bigl( \bigl(m^{\alpha}_{\mu}\bigr)^2-\bigl(m^{\alpha}_{\nu} \bigr)^2\bigr) \bigl(t^{(k)}\bigr)dt \biggr| \\ &\quad{} \leq\exp\biggl\{\frac{k_{\alpha}T}{2\lambda_{\alpha}^2} \biggr\} \biggl|\int\exp\biggl( - \frac {1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2dt \biggr) d(\gamma_{K_{\mu }}-\gamma_{K_{\nu}}) \biggr|. \end{aligned}$$

Let ξ be a probability measure on \(\mathcal{C}\times\mathcal {C}\) with marginals μ and ν, and let γ ξ be the law of a bidimensional centered Gaussian process \((G^{\alpha}, \widetilde{G^{\alpha}})\) with covariance \(K^{\alpha}_{\xi}\):

$$ \begin{aligned}[b] &K^{\alpha}_{\xi}(s,t)\\ &\quad{} = \sum_{\gamma=1}^M \frac{\sigma_{\alpha \gamma}^2}{\lambda_{\alpha}^2} \left ( \begin{array}{c@{\quad}c} \int S_{\alpha\gamma}(x^{\gamma}_{s-\tau_{\alpha\gamma }})S_{\alpha\gamma}(x^{\gamma}_{t-\tau_{\alpha\gamma}}) \, d\xi(x,y) & \int S_{\alpha\gamma}(x^{\gamma}_{s-\tau_{\alpha\gamma }})S_{\alpha\gamma}(y^{\gamma}_{t-\tau_{\alpha\gamma}}) \, d\xi (x,y) \\ \int S_{\alpha\gamma}(y^{\gamma}_{s-\tau_{\alpha\gamma }})S_{\alpha\gamma}(x^{\gamma}_{t-\tau_{\alpha\gamma}}) \, d\xi(x,y) & \int S_{\alpha\gamma}(y^{\gamma}_{s-\tau_{\alpha\gamma }})S_{\alpha\gamma}(y^{\gamma}_{t-\tau_{\alpha\gamma}}) \, d\xi (x,y) \\ \end{array} \right ) . \end{aligned} $$
(34)

Then,

$$\begin{aligned} &\biggl|\int\exp\biggl( - \frac{1}{2} \int_0^T {G^{\alpha}_{t^{(k)} }}^2dt \biggr) d( \gamma_{K_{\mu}}-\gamma_{K_{\nu}}) \biggr|\\ &\quad{} = \biggl|\int\biggl\{ \exp\biggl( - \frac{1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2dt \biggr) - \exp\biggl( - \frac {1}{2} \int_0^T \widetilde{G^{\alpha}_{t^{(k)}}}^2dt \biggr) \biggr\} d \gamma_{\xi} \biggr| \\ &\quad{} \leq\frac{1}{2} \int\int_0^T \bigl| {G^{\alpha}_{t^{(k)} }}^2-\widetilde{G^{\alpha}_{t^{(k)}}}^2 \bigr| dt d\gamma_{\xi} \\ &\quad{} \leq\frac{1}{2} \prod_{\varepsilon= \pm1} \biggl(\int\int _0^T \bigl(G^{\alpha }_{t^{(k)}}+ \varepsilon\widetilde{G^{\alpha}_{t^{(k)}}}\bigr)^2 dt d \gamma_{\xi} \biggr)^{\frac{1}{2}} \end{aligned}$$

by Cauchy-Schwarz inequality. Then, using the covariance of \((G^{\alpha },\widetilde{G^{\alpha}})\) under γ ξ , we find:

$$\begin{aligned} & \biggl|\varGamma_1^{\alpha,k}(\mu)-\varGamma_1^{\alpha,k}( \nu)+\frac{1}{2}\int\bigl(\bigl(m^{\alpha}_{\mu} \bigr)^2-\bigl(m^{\alpha}_{\nu}\bigr)^2 \bigr) \bigl(t^{(k)}\bigr)dt \biggr| \\ & \quad{}\leq\frac{1}{2} \exp\biggl\{\frac{k_{\alpha}T}{2\lambda _{\alpha}^2} \biggr\} \biggl( \frac{4k_{\alpha}T}{\lambda_{\alpha}^2} \biggr)^{\frac{1}{2}} \Biggl\{ \frac{1}{\lambda_{\alpha} ^2}\sum _{\gamma=1}^M \sigma_{\alpha\gamma}^2 \int\int_0^T \bigl(S_{\alpha \gamma} \bigl(x^{\gamma}_{t-\tau_{\alpha\gamma}}\bigr) \\ & \qquad{}-S_{\alpha\gamma} \bigl(y^{\gamma}_{t-\tau_{\alpha\gamma}}\bigr)\bigr)^2 dt \, d\xi(x,y) \Biggr\} ^{\frac{1}{2}} \\ & \quad{}\leq\frac{k_{\alpha}}{\lambda_{\alpha}^2} \sqrt{T} \exp\biggl\{ \frac{k_{\alpha} T}{2\lambda_{\alpha} ^2} \biggr \} \max_{\gamma=1\cdots M} \biggl\{ \int\int_0^T \bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t-\tau_{\alpha\gamma}}\bigr )-S_{\alpha\gamma} \bigl(y^{\gamma }_{t-\tau_{\alpha\gamma}}\bigr) \bigr|^2 dt \, d\xi(x,y) \biggr\}^{\frac{1}{2}}. \end{aligned}$$
(35)

Moreover, we have:

$$\begin{aligned} \biggl|\int{m^{\alpha}_{\mu}\bigl(t^{(k)} \bigr)}^2 -{m^{\alpha}_{\nu}\bigl(t^{(k)} \bigr)}^2 dt \biggr| = &\int\bigl| \bigl(m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr) -m^{\alpha}_{\nu} \bigl(t^{(k)} \bigr) \bigr) \bigl(m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr) + m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr) \bigr) \bigr| dt \\ \leq&2 \frac{\bar{J}_{\alpha}}{\lambda_{\alpha}} \int\bigl|m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr) -m^{\alpha }_{\nu} \bigl(t^{(k)}\bigr) \bigr| dt. \end{aligned}$$

But

$$\begin{aligned} \int_0^T \bigl|\bigl(m^{\alpha}_{\mu}-m^{\alpha}_{\nu} \bigr) (t)\bigr| dt = &\int_0^T \Biggl|\frac{1}{\lambda_{\alpha}} \sum_{\gamma=1}^M \bar{J}_{\alpha \gamma}\int S_{\alpha\gamma}\bigl(x^{\gamma }_{t-{\tau}}\bigr) d(\mu-\nu) (x) \Biggr| \, dt \\ \leq&\frac{1}{\lambda_{\alpha}} \sum_{\gamma=1}^M \bigl| \bar{J}_{\alpha\gamma}\bigr| \int_0^T \biggl|\int S_{\alpha\gamma}\bigl(x^{\gamma}_{t-{\tau_{\alpha\gamma}}}\bigr) d(\mu -\nu) (x) \biggr| \, dt \\ \leq&\frac{1}{\lambda_{\alpha}} \sum_{\gamma=1}^M | \bar{J}_{\alpha\gamma}| \int\int_0^T \bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t-{\tau_{\alpha\gamma}}}\bigr) - S_{\alpha\gamma }\bigl(y^{\gamma}_{t-{\tau}}\bigr)\bigr| dt \, d\xi(x,y) \\ \leq&\frac{ \bar{J}_{\alpha}}{\lambda_{\alpha}} \max_{\gamma =1\cdots M} \biggl( \int\int _0^T \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_{t-{\tau_{\alpha\gamma}}} \bigr) - S_{\alpha\gamma}\bigl(y^{\gamma}_{t-{\tau }}\bigr) \bigr|^2 dt \, d\xi(x,y) \biggr)^{\frac{1}{2}} \end{aligned}$$

by Cauchy-Schwarz inequality.

Consequently,

$$\begin{aligned} \bigl|\varGamma_1^{\alpha,k}(\mu) - \varGamma_1^{\alpha,k}( \nu)\bigr| \leq&\biggl(\frac{ \bar{J}_{\alpha}^2}{\lambda_{\alpha}^2} + \frac{k_{\alpha }}{\lambda_{\alpha}^2} \sqrt{T} \exp\biggl\{ \frac{k_{\alpha} T}{2\lambda_{\alpha} ^2} \biggr\} \biggr) \max_{\gamma=1\cdots M} \biggl( \int \int_0^T \bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}} \bigr) \\ &{} - S_{\alpha\gamma }\bigl(y^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr) \bigr|^2 dt \, d\xi(x,y) \biggr)^{\frac{1}{2}} . \end{aligned}$$
(36)

As the S αγ are K S Lipschitz, we have:

$$\begin{aligned} \bigl|\varGamma_1^{\alpha,k}(\mu) - \varGamma_1^{\alpha,k}( \nu)\bigr| \leq K_S \sqrt{T} \biggl(\frac{\bar{J}_{\alpha}^2}{\lambda _{\alpha}^2} + \frac {k_{\alpha}}{\lambda_{\alpha}^2} \sqrt{T} \exp\biggl\{ \frac{k_{\alpha }T}{2\lambda_{\alpha}^2} \biggr\} \biggr) d_T(\mu,\nu), \end{aligned}$$
(37)

so that \(\varGamma_{1}^{\alpha,k}\) is Lipschitz for the Vaserstein distance. Using the triangle inequality, the result holds for \(\varGamma_{1}^{k}\).

Proof of Lemma 5.(ii)

Let

$$\begin{aligned} F_{\mu}(x) =& \log\biggl\{ \int\exp\biggl\{ \int_0^T \bigl(\mathbf{G}_{t^{(k)} }(\omega)+\mathbf{m}_{\mu} \bigl(t^{(k)}\bigr) \bigr)' \cdot d\mathbf{W}_t(x) \\ &{}- \frac{1}{2} \int_0^T \bigl\| \mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\mu } \bigl(t^{(k)}\bigr) \bigr\|^2 dt \biggr\} d\gamma_{\mu} \biggr\}. \end{aligned}$$

This function is a.s. finite but not bounded, let us hence define for \(A \in\mathbb{R}^{+}\)

$$\begin{aligned} F_{\mu}^A(x) =& \log\biggl\{ \int A\wedge\exp\biggl\{ \int_0^T \bigl(\mathbf{G}_{t^{(k)} }( \omega)+\mathbf{m}_{\mu}\bigl(t^{(k)}\bigr) \bigr)' \cdot d\mathbf{W}_t(x) \\ &{}- \frac{1}{2} \int_0^T \bigl\| \mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\mu } \bigl(t^{(k)}\bigr) \bigr\|^2 dt \biggr\} d\gamma_{\mu} \biggr\}. \end{aligned}$$

By the monotone convergence theorem and using Eq. (22), we have for any a≥1:

$$\begin{aligned} a \int F_{\mu}(x) d\mu(x) \leq I(\mu|P)+ \log\biggl\{ \int\exp{ aF_{\mu }(x)} dP(x) \biggr\}. \end{aligned}$$

By Jensen inequality and Fubini theorem,

$$\begin{aligned} \int\exp{ \bigl(aF_{\mu}(x)}\bigr) dP(x) \leq&\int\prod _{\alpha=1}^M \int\exp\biggl\{ a \int _0^T \bigl(G^{\alpha}_{t^{(k)}} + m^{\alpha}_{\mu }\bigl(t^{(k)}\bigr) \bigr) dW^{\alpha}_t(x) \biggr\} dP_{\alpha}(x) \\ &{}\times \exp\biggl\{ - \frac{a}{2} \int_0^T \bigl\|\mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\mu} \bigl(t^{(k)} \bigr) \bigr\|^2 dt \biggr\} d\gamma_{\mu}. \end{aligned}$$

But, as W α is a P α -Brownian motion,

$$\begin{aligned} \int\exp\biggl\{ a \int_0^T \bigl(G^{\alpha}_{t^{(k)}} + m^{\alpha }_{\mu} \bigl(t^{(k)} \bigr) \bigr) dW^{\alpha}_t(x) \biggr\} dP_{\alpha}(x) = \exp\biggl\{ \frac {a^2}{2} \int _0^T \bigl(G^{\alpha}_{t^{(k)}} + m^{\alpha}_{\mu}\bigl(t^{(k)}\bigr) \bigr)^2 dt \biggr\} , \end{aligned}$$

so that

$$\begin{aligned} a \int F_{\mu}(x) d\mu(x) \leq I(\mu|P)+ \log\biggl\{ \int\exp\biggl \{ \frac{a^2-a}{2} \int_0^T \bigl\| \mathbf{G}_{t^{(k)}}(\omega)+\mathbf{m}_{\mu } \bigl(t^{(k)}\bigr)\bigr \|^2 dt \biggr\} d\gamma_{\mu} \biggr\}. \end{aligned}$$

Letting a=1 proves that Γ kI(|P).

Proof of Lemma 5.(iii)

As the components of G are independent under γ μ , we only have to check that, for every b>0, there exists a finite constant C b such that

$$\begin{aligned} {\mathcal{E}}_{\mu} \biggl[ \exp{ \biggl(\frac{b}{2} \int _0^T \bigl(G^{\alpha }_s+m^{\alpha}_{\mu}(s) \bigr)^2 ds \biggr) } \biggr] \leq\exp{\frac {b C_b k_{\alpha}T}{\lambda_{\alpha}^2}} . \end{aligned}$$
(38)

It was proved in [3, Lemma A.3(2)] in their particular framework that for every b verifying \(\frac{bk_{\alpha }T}{\lambda_{\alpha} ^{2}}<1\), there exists a finite constant c b such that:

$$\begin{aligned} \int\exp{ \biggl(\frac{b}{2} \int_0^T G_s^2 ds \biggr) } d\gamma_{\mu } \leq\exp{ \frac{b c_b k_{\alpha}T}{\lambda_{\alpha}^2}}. \end{aligned}$$

In our case, the covariance function is slightly different of that of [3], but the proof and result remain unchanged and can be readily extended.

Moreover, since we have

$$\begin{aligned} \bigl(G^{\alpha}_s + m^{\alpha}_{\mu}(s) \bigr)^2 \leq2 {G^{\alpha}_s}^2 + 2 {m^{\alpha}_{\mu}(s)}^2 \leq2{G^{\alpha}_s}^2 + 2\frac{\bar {J}_{\alpha} ^2}{\lambda_{\alpha}^2} \end{aligned}$$

we obtain the desired result with the following constant \(C_{b} = 2 c_{2b} + \frac{\bar{J}_{\alpha}^{2}}{k_{\alpha}}\), under the condition \(\frac{2b k_{\alpha} T}{\lambda_{\alpha} ^{2}}<1\).

Proof of Lemma 5.(iv)

As above, let us prove the result for \(| \varGamma^{\alpha ,k}_{2,\nu} - \varGamma^{\alpha,k}_{2} |\). We have:

$$\begin{aligned} \bigl|\varGamma_{2,\nu}^{\alpha,k}(\mu) - \varGamma_2^{\alpha,k}( \mu)\bigr| \leq&\frac {1}{2} \biggl|\int\int\biggl\{ \biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha }_t - m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 \\ &{}- \biggl( \int G^{\alpha }_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu}\bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 \biggr\} d\gamma_{\widetilde{K}_{\mu}^{T,k}} d\mu\biggr| \\ &{}+\frac{1}{2} \biggl|\int\int\biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu } \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d ( \gamma_{\widetilde{K}_{\nu}^{T,k}}-\gamma_{\widetilde {K}_{\mu}^{T,k}} ) d\mu\biggr|\\ &{} + \biggl|\int\int \bigl(m^{\alpha}_{\nu }-m^{\alpha }_{\mu}\bigr) \bigl(t^{(k)}\bigr)dW^{\alpha}_t d\mu\biggr|. \end{aligned}$$

Let ξ be a probability measure on \(\mathcal{C}\times\mathcal {C}\) with marginals μ and ν, and let γ ξ be the law of a bidimensional centered Gaussian process \((G^{\alpha}, \widetilde{G^{\alpha}})\) with covariance \(K^{\alpha}_{\xi}\). Let

$$\begin{aligned} \varLambda_T^{\alpha,k}\bigl(G^{\alpha}\bigr)= \frac{\exp{ ( -\frac {1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2 dt )}}{\int\exp{ ( -\frac {1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2 dt )} \, d\gamma_{\xi}}. \end{aligned}$$

As in [3, Lemma 3.4], we can show that:

$$ \begin{aligned}[b] &\bigl|\varGamma_{2,\nu}^{\alpha,k}(\mu) - \varGamma_2^{\alpha,k}( \mu)\bigr|\\ &\quad{} \leq\frac {1}{2} \overbrace{\int\int\bigl|\varLambda_T^{\alpha,k} \bigl(G^{\alpha }\bigr)-\varLambda_T^{\alpha,k}\bigl( \widetilde{G^{\alpha}}\bigr) \bigr| \biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\gamma_{\xi} d\mu}^{B_1} \\ &\qquad{}+ \underbrace{\frac{1}{2} \! \prod_{\varepsilon=\pm1} \biggl( \int\! \int\varLambda_T^{\alpha ,k}\bigl(\widetilde {G^{\alpha}}\bigr) \biggl( \int\bigl(G^{\alpha}_{t^{(k)}} + \varepsilon\widetilde{G^{\alpha}}_{t^{(k)}}\bigr) \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\gamma_{\xi} d\mu\biggr)^{\frac{1}{2}}}_{B_2} \\ &\qquad{}+ \frac{1}{2} \underbrace{ \biggl|\int\!\!\int\varLambda_T^{\alpha,k} \bigl(G^{\alpha}\bigr) \biggl\{ \biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 - \biggl( \int G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu }\bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 \biggr\} d\gamma_{\xi} d\mu\biggr|}_{B_3} \\ &\qquad{}+ \underbrace{ \biggl( \int\biggl| \int\bigl(m^{\alpha}_{\nu}-m^{\alpha}_{\mu} \bigr) \bigl(t^{(k)} \bigr)dW^{\alpha}_t \biggr|^2 d\mu\biggr)^{\frac{1}{2}}}_{B_4} \end{aligned} $$
(39)

and

$$\begin{aligned} \varLambda_T^{\alpha,k}\bigl(G^{\alpha}\bigr)= \exp\biggl \{ - \varGamma_1^{\alpha,k}(\mu) - \frac{1}{2} \int _0^T {m^{\alpha}_{\mu}}^2 \bigl(t^{(k)}\bigr) dt - \frac {1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2 dt \biggr\}. \end{aligned}$$

Hence, we have by Jensen inequality,

$$\begin{aligned} \varLambda_T^{\alpha,k}\bigl(G^{\alpha}\bigr) \leq\exp \biggl\{\frac {k_{\alpha} T}{2\lambda_{\alpha}^2} \biggr\}, \end{aligned}$$

so that

$$\begin{aligned} \bigl|\varLambda_T^{\alpha,k}\bigl(G^{\alpha}\bigr)- \varLambda_T^{\alpha ,k}\bigl(\widetilde{G^{\alpha}}\bigr) \bigr| \leq&\exp\biggl\{\frac{k_{\alpha}T}{2\lambda _{\alpha}^2} \biggr\} \biggl(\frac {1}{2}\int _0^T \bigl| {G^{\alpha}_{t^{(k)}}}^2- \widetilde{G^{\alpha }}_{t^{(k)} }^2 \bigr|dt\\ &{} + \biggl| \varGamma_1^{\alpha,k}(\mu)-\varGamma_1^{\alpha ,k}( \nu)+\frac{1}{2}\int\bigl(\bigl(m^{\alpha}_{\mu} \bigr)^2-\bigl(m^{\alpha}_{\nu}\bigr)^2 \bigr) \bigl(t^{(k)} \bigr)dt \biggr| \biggr), \end{aligned}$$

which eventually gives

$$\begin{aligned} B_1 \leq&\frac{1}{2}\exp\biggl\{\frac{k_{\alpha}T}{2\lambda_{\alpha }^2} \biggr\} \biggl( \biggl|\varGamma_1^{\alpha,k}(\mu)-\varGamma_1^{\alpha,k}( \nu)+\frac {1}{2}\int\bigl(\bigl(m^{\alpha}_{\mu} \bigr)^2-\bigl(m^{\alpha}_{\nu}\bigr)^2 \bigr) \bigl(t^{(k)}\bigr)dt \biggr|\\ &{}\times \int\int\biggl( \int _0^T G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu} \bigl(t^{(k)} \bigr)dt\bigr) \biggr)^2 d\gamma_{\xi}d \mu \\ &{}+ \int\int\biggl(\int_0^T \bigl| \bigl(G^{\alpha}_{t^{(k)}}\bigr)^2-\widetilde {G^{\alpha }}_{t^{(k)}}^2 \bigr| dt \biggr) \biggl( \int _0^T G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t - m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\gamma_{\xi}d \mu\biggr). \end{aligned}$$

Let h,mL 2([0;T],dt), with m bounded. By Cauchy-Schwarz and the relative entropy inequality (22) (with \(\varPhi(x)= ( \int_{0}^{T} h_{t} dW^{\alpha}_{t}(x) )^{2} \sim \mathcal {N} (0,\int_{0}^{T} h_{t}^{2} dt )^{2}\) under P α , and besides is a positive and measurable function of \(\mathcal{C}\)), we have the existence of a finite constant C such that,

$$\begin{aligned} \int\biggl( \int_0^T h_t \bigl(dW^{\alpha}_t(x) - m(t)dt\bigr) \biggr)^2 d\mu (x) \leq&2 \biggl\{ \int\biggl(\int_0^T h_t dW^{\alpha}_t \biggr)^2 + \biggl( \int_0^T h_t m_t dt \biggr)^2 d\mu\biggr\} \\ \leq&2\biggl \{ \bigl(C \bigl(1+I(\mu|P) \bigr) + m^2_{\infty}T \bigr) \biggl( \int_0^T h_t^2 dt \biggr)\biggr \} \\ \leq &C' \bigl(1+I(\mu|P) \bigr) \biggl( \int _0^T h_t^2 dt \biggr) . \end{aligned}$$
(40)

We can now bound the different terms in inequality (39).

In fact, as \(h_{t}=G^{\alpha}_{t^{(k)}}\) and \(m_{t}=m^{\alpha}_{\nu}(t^{(k)})\) verify the required condition, (40) gives the existence of c T ,

$$\begin{aligned} \int\biggl( \int_0^T G^{\alpha}_{t^{(k)}} \bigl(dW^{\alpha}_t(x) - m^{\alpha }_{\nu } \bigl(t^{(k)}\bigr)dt\bigr) \biggr)^2 d\mu(x) \leq\: c_T \bigl(1+I(\mu|P) \bigr) \int_0^T \bigl(G^{\alpha}_{t^{(k)}}\bigr)^2 dt. \end{aligned}$$

Hence, we can find a finite constant \(c'_{T}\) such that

$$\begin{aligned} B_1 \leq c'_T \bigl(1+I(\mu|P) \bigr) \max_{\gamma=1\cdots M} \biggl( \int\int_0^T \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha \gamma}}\bigr)-S_{\alpha\gamma} \bigl(y^{\gamma}_{t^{(k)} -\tau_{\alpha\gamma} }\bigr) \bigr|^2 dt \, d\xi(x,y) \biggr)^{\frac{1}{2}}. \end{aligned}$$

Similarly, there exists a constant c T such that

$$\begin{aligned} B_2 \leq&\frac{1}{2}\exp\biggl\{\frac{k_{\alpha}T}{2\lambda _{\alpha}^2} \biggr \} \prod_{\varepsilon=\pm1} \biggl( c_T \bigl(1+I( \mu|P) \bigr) \int\int\bigl(G^{\alpha }_{t^{(k)}} +\varepsilon \widetilde{G^{\alpha}}_{t^{(k)}}\bigr)^2 dt d\gamma _{\xi} \biggr)^{\frac{1}{2}} \\ \leq &c'_T \bigl(1+I(\mu|P) \bigr) \max _{\gamma=1\cdots M} \biggl( \int\int_0^T \bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma }}\bigr)-S_{\alpha\gamma} \bigl(y^{\gamma}_{t^{(k)} -\tau_{\alpha\gamma}}\bigr) \bigr|^2 dt \, d\xi(x,y) \biggr)^{\frac{1}{2}}. \end{aligned}$$

To bound B 3, we first use Cauchy-Schwarz inequality:

$$\begin{aligned} B_3 \leq&\frac{1}{2} \exp\biggl\{ \frac{k_{\alpha}T}{2\lambda _{\alpha}^2} \biggr\} \prod_{\varepsilon= \pm1} \biggl\{ \int\int\biggl| \int _0^T G^{\alpha }_{t^{(k)} } \bigl( (1+ \varepsilon)dW^{\alpha}_t \\ &{}- \bigl(m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr) + \varepsilon m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr)\bigr)dt \bigr)\biggr |^2 d\gamma _{\xi} d\mu\biggr\}^{\frac{1}{2}} . \end{aligned}$$
(41)

But

$$\begin{aligned} \biggl| \int_0^T G^{\alpha}_{t^{(k)}} \bigl(m^{\alpha}_{\mu}\bigl(t^{(k)}\bigr) -m^{\alpha }_{\nu}\bigl(t^{(k)}\bigr) \bigr) dt \biggr|^2 \leq\biggl(\int_0^T {G_{t^{(k)} }^{\alpha}}^2 dt \biggr) \biggl( \int _0^T \bigl(m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr) -m^{\alpha }_{\nu} \bigl(t^{(k)} \bigr) \bigr)^2 dt \biggr). \end{aligned}$$

Remark that

$$\begin{aligned} \int_0^T \bigl(m^{\alpha}_{\mu} \bigl(t^{(k)}\bigr) - m^{\alpha}_{\nu } \bigl(t^{(k)}\bigr) \bigr)^2 dt = &\int _{0}^T \Biggl( \sum _{\gamma=1}^M \frac{\bar {J}_{\alpha\gamma}}{\lambda_{\alpha}} \int S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr) d(\mu- \nu) (x) \Biggr)^2 dt \\ \leq&\frac{\bar{J}_{\alpha}^2}{\lambda_{\alpha}^2} \max_{\gamma =1\cdots M} \int _{0}^T \biggl( \int S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr) d(\mu- \nu) (x) \biggr)^2 dt \\ \leq&\frac{\bar{J}_{\alpha}^2}{\lambda_{\alpha}^2} \max_{\gamma =1\cdots M} \int _{0}^T \biggl( \int\bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\\ &{} - S_{\alpha\gamma} \bigl(y^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma} }\bigr)\bigr| d\xi(x,y) \biggr)^2 dt. \end{aligned}$$

So that

$$\begin{aligned} &\biggl| \int_0^T G^{\alpha}_{t^{(k)}} \bigl(m^{\alpha}_{\mu}\bigl(t^{(k)}\bigr) -m^{\alpha }_{\nu}\bigl(t^{(k)}\bigr) \bigr) dt \biggr|^2 \\ &\quad{}\leq\frac{\bar{J}_{\alpha}^2}{\lambda_{\alpha}^2} \biggl(\int_0^T G^2_{t^{(k)}} dt \biggr) \max_{\gamma =1\cdots M} \int _{0}^T \biggl( \int\bigl|S_{\alpha\gamma} \bigl(x^{\gamma }_{t^{(k)}-\tau_{\alpha\gamma}}\bigr) - S_{\alpha\gamma} \bigl(y^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\bigr| d\xi(x,y) \biggr )^2 dt. \end{aligned}$$

Moreover, (40) gives:

$$\begin{aligned} \int\biggl\{ \int_0^T 2 G^{\alpha}_{t^{(k)}} \biggl(dW^{\alpha}_t - \frac {m^{\alpha}_{\mu}(t^{(k)})+m^{\alpha}_{\nu}(t^{(k)})}{2} dt \biggr) \biggr\} ^2 d\mu \leq c_T \bigl( 1+ I(\mu|P) \bigr) 4 \int _0^T {G^{\alpha}_{t^{(k)} }}^2 dt. \end{aligned}$$

Using the last two inequalities in (41) we have:

$$\begin{aligned} B_3 \leq&\frac{1}{2} \exp\biggl\{ \frac{k_{\alpha}T}{2\lambda _{\alpha}^2} \biggr\} \biggl\{ \int c_T \bigl( 1+ I(\mu|P) \bigr) 4 \biggl(\int _0^T {G^{\alpha}_{t^{(k)}}}^2 dt \biggr) d\gamma_{\xi} \biggr\}^{\frac{1}{2}} \\ &{}\times\biggl\{ \int \frac{\bar{J}_{\alpha}^2}{\lambda_{\alpha}^2} \biggl(\int_0^T {G^{\alpha}_{t^{(k)}}}^2 dt \biggr) \max _{\gamma=1\cdots M} \int_{0}^T \biggl( \int\bigl|S_{\alpha \gamma} \bigl(x^{\gamma }_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\\ &{} -S_{\alpha\gamma}\bigl(y^{\gamma }_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\bigr| d\xi(x,y) \biggr)^2 dt d\gamma_{\xi} \biggr\}^{\frac{1}{2}} \\ \leq &c'_T \bigl( 1+ I(\mu|P) \bigr) \max _{\gamma=1\cdots M} \biggl\{\int_{0}^T \int\bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha \gamma}}\bigr) - S_{\alpha\gamma}\bigl(y^{\gamma}_{t^{(k)} -\tau_{\alpha\gamma} }\bigr)\bigr|^2 d \xi(x,y) dt \biggr\}^{\frac{1}{2}} \end{aligned}$$

as I(|P)≥0.

As of the last term, we have

$$\begin{aligned} B_4 \leq&\biggl( c_T \bigl(1+I(\mu|P) \bigr) \int _0^T \bigl(m^{\alpha}_{\mu } \bigl(t^{(k)}\bigr) - m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr) \bigr)^2 dt \biggr)^{\frac {1}{2}} \\ \leq&\frac{\bar{J}_{\alpha}}{\lambda_{\alpha}} \bigl( c_T \bigl(1+I(\mu |P) \bigr) \bigr)^{\frac {1}{2}} \max_{\gamma=1\cdots M} \biggl(\int _{0}^T \biggl( \int\bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\\ &{} - S_{\alpha\gamma } \bigl(y^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\bigr| d\xi(x,y) \biggr )^2 dt \biggr)^{\frac{1}{2}} \\ \leq &c'_T \bigl(1+I(\mu|P) \bigr) \max _{\gamma=1\cdots M} \biggl(\int_{0}^T \int\bigl|S_{\alpha\gamma}\bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha \gamma}}\bigr) - S_{\alpha\gamma}\bigl(y^{\gamma}_{t^{(k)} -\tau_{\alpha\gamma} }\bigr)\bigr|^2 d \xi(x,y) dt \biggr)^{\frac{1}{2}}. \end{aligned}$$

We have proved that there exist a constant c T such that

$$ \begin{aligned}[b] &\bigl|\varGamma_{2,\nu}^{\alpha,k}(\mu) - \varGamma_2^{\alpha,k}( \mu)\bigr|\\ &\quad{}\leq c_T \bigl(1+I(\mu|P) \bigr) \max_{\gamma=1\cdots M} \biggl(\int_{0}^T \int\bigl|S_{\alpha\gamma} \bigl(x^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr) - S_{\alpha\gamma } \bigl(y^{\gamma}_{t^{(k)}-\tau_{\alpha\gamma}}\bigr)\bigr|^2 d\xi(x,y) dt \biggr)^{\frac{1}{2}}. \end{aligned} $$
(42)

Therefore

$$\begin{aligned} \bigl|\varGamma_{2,\nu}^{\alpha,k}(\mu) - \varGamma_2^{\alpha,k}( \mu)\bigr| \leq c_T K_S \sqrt{T} \bigl(1+I(\mu|P) \bigr) d_T(\mu,\nu), \end{aligned}$$

so that using the triangle inequality

$$\begin{aligned} \bigl|\varGamma_{2,\nu}^{k}(\mu) - \varGamma_2^{k}( \mu)\bigr| \leq C_T \bigl(1+I(\mu|P) \bigr) d_T(\mu,\nu). \end{aligned}$$

Proof of Lemma 5.(v)

For all α∈{1⋯M}, let

$$\begin{aligned} dQ_{\nu}^{\alpha, k}(x) = \exp{\varGamma_{\nu}^{\alpha ,k}( \delta_x)}dP_{\alpha} (x) = &\int\exp \biggl( \int_0^T \bigl(G^{\alpha}_{t^{(k)}} + m^{\alpha }_{\nu} \bigl(t^{(k)} \bigr)\bigr) \, dW^{\alpha}_t(x) \\ &{}- \frac{1}{2} \int_0^T \bigl(G^{\alpha}_{t^{(k)}} + m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr)\bigr)^2 dt \biggr) \, d \gamma_{\nu} \, dP_{\alpha}(x). \end{aligned}$$

The equality between the two expression of \(Q_{\nu}^{\alpha,k}\) is easily obtained by Gaussian calculus (see the proof of Proposition 4). We deduce by the martingale property of this density that it is a probability measure on \(\mathcal{C}([-\tau ,T],\mathbb{R})\).

Remark that

It follows that \(Q_{\nu}^{k} \in\mathcal{M}_{1}^{+}(\mathcal{C})\), and

$$\begin{aligned} dQ_{\nu}^k(x) = \exp{\varGamma_{\nu}^{k}( \delta_x)}dP(x) = &\int\exp \biggl( \int_0^T \bigl( \mathbf{G}_{t^{(k)}} + \mathbf{m}_{\nu}\bigl(t^{(k)} \bigr) \bigr)' \cdot\, d\mathbf{W}_t(x) \\ &{} - \frac{1}{2} \int_0^T \bigl\| \mathbf {G}_{t^{(k)}} + \mathbf{m}_{\nu}\bigl(t^{(k)} \bigr) \bigr\|^2 dt \biggr) \, d\gamma_{\nu} \, dP(x). \end{aligned}$$

Lets now prove that \(H^{k}_{\nu} =I(.|Q^{k}_{\nu})\). We will first show that \(I(Q^{k}_{\nu}|P)\) is finite. In fact,

$$\begin{aligned} \frac {\mbox {d} Q_{\nu}^{\alpha, k}}{\mbox {d} P_{\alpha}} (x) = &\biggl(\int\exp \biggl\{ -\frac{1}{2} \int_0^T {G^{\alpha}_{t^{(k)}}}^2 + {m^{\alpha}_{\nu}}^2 \bigl(t^{(k)}\bigr) dt \biggr\} d\gamma_{\nu} \biggr) \exp \biggl\{\int_0^T m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr) dW^{\alpha }_t(x) \biggr\} \\ &{}\times\exp\biggl\{\frac{1}{2} \int\biggl(\int_0^T G^{\alpha }_{t^{(k)}} \bigl( dW^{\alpha}_t(x) - m^{\alpha}_{\nu}\bigl(t^{(k)}\bigr) dt \bigr) \biggr)^2 d\gamma_{\widetilde{K}_{\nu}^{T,k}} \biggr\}. \end{aligned}$$
(43)

Which becomes, after some Gaussian computations (see [3, Lemma 5.15]):

$$\begin{aligned} \frac {\mbox {d} Q_{\nu}^{\alpha, k}}{\mbox {d} P_{\alpha}} = \exp\biggl\{\int _0^T H^{\alpha }_{t^{(k)} } \bigl(Q_{\nu}^{\alpha, k}\bigr)dW^{\alpha}_t- \frac{1}{2}\int_0^T {H^{\alpha }_{t^{(k)} }}^2 \bigl(Q_{\nu}^{\alpha, k}\bigr)dt \biggr\} \end{aligned}$$

where

$$\begin{aligned} H^{\alpha}_{t^{(k)}}\bigl(Q^{\alpha, k}_{\nu}\bigr) = & \biggl(\int G^{\alpha }_{t^{(k)}} \int_0^t G^{\alpha}_{s^{(k)}} \bigl(dW^{\alpha}_s - m^{\alpha}_{\nu }\bigl(s^{(k)} \bigr)ds \bigr) \; d \gamma_{\widetilde{K}_{\nu}^{t,k}} \biggr) + m^{\alpha }_{\nu } \bigl(t^{(k)}\bigr) \\ = &\int_0^t \bigl(\widetilde{K}_{\nu}^{t,k} \bigl(t^{(k)},s^{(k)}\bigr) \bigr)_{\alpha\alpha } \bigl(dW^{\alpha}_s - m^{\alpha}_{\nu} \bigl(s^{(k)}\bigr)ds \bigr) + m^{\alpha }_{\nu } \bigl(t^{(k)}\bigr) \\ = &\sum_{l=0,\ldots,k ; t_{l+1}\leq t} \bigl(W^{\alpha }_{t_{l+1}}-W^{\alpha }_{t_l} - m^{\alpha}_{\nu}(t_l) (t_{l+1}-t_l) \bigr) \bigl(\widetilde{K}_{\nu}^{t,k}\bigl(t^{(k)}, t_k\bigr) \bigr)_{\alpha\alpha} + m^{\alpha }_{\nu } \bigl(t^{(k)}\bigr). \end{aligned}$$

Hence, according to Girsanov Theorem, there exists a \(Q_{\nu}^{\alpha, k}\)-Brownian motion B α such that

$$\begin{aligned} W^{\alpha}_t=B^{\alpha}_t+\int _0^t H^{\alpha}_{s^{(k)}} \bigl(Q_{\nu }^{\alpha, k}\bigr)ds. \end{aligned}$$

As W is an affine function of B α, it is, under \(Q_{\nu }^{\alpha, k}\), a Gaussian variable with finite moments. In particular, \(I(Q_{\nu}^{\alpha, k}|P_{\alpha})\) is finite. But

$$\begin{aligned} I\bigl(Q_{\nu}^k|P\bigr)= \int_{\mathcal{C}} \log\biggl( \frac {\mbox {d} Q^k_{\nu }}{\mbox {d} P} (x) \biggr) dQ^k_{\nu}(x) =& \sum_{\alpha=1}^M \int_{\mathcal{C}([-\tau,T],\mathbb{R})} \log\biggl( \frac {\mbox {d} Q^{\alpha, k}_{\nu}}{\mbox {d} P_{\alpha}} \bigl(x^{\alpha}\bigr) \biggr) dQ^{\alpha, k}_{\nu } \bigl(x^{\alpha }\bigr) \\ =& \sum_{\alpha=1}^M I\bigl(Q_{\nu}^{\alpha, k}|P_{\alpha}\bigr), \end{aligned}$$

so that \(I(Q_{\nu}^{k}|P)\) is finite.

Now, let

$$\begin{aligned} \biggl(\tau^{\alpha}_{m}(x)=\inf\biggl\{t \geq0 ; \biggl| \int _0^t m^{\alpha}_{\nu} \bigl(s^{(k)}\bigr) dW^{\alpha}_s(x) \biggr| \geq m \biggr\} \biggr)_{m \in \mathbb{N}} \end{aligned}$$

be a sequence of stopping times for the Brownian filtration \(\sigma( W^{\alpha}_{s}, 0\leq s \leq T)\). As W α is a P α -MB, we have \(\lim_{m\to\infty} \tau ^{\alpha }_{m} \to\infty\) almost surely under P α . \(( \tau_{m}= \min_{\alpha=1 \cdots M} \tau^{\alpha}_{m} )_{m\in\mathbb{N} }\) defines a sequence stopping times for \(\sigma( W^{\alpha}_{s}, \alpha =1\cdots M, 0\leq s \leq T)\) which tends to infinity along with m P a.s. We define:

$$\begin{aligned} Q_{\nu,t}^k=\int\exp\biggl\{ \biggl( \int _0^t \bigl(\mathbf{G}_{s^{(k)}} + \mathbf{m}_{\nu}\bigl(s^{(k)} \bigr)\bigr)' \cdot \, d\mathbf{W}_s - \frac{1}{2} \int_0^t \bigl\|\mathbf{G}_{s^{(k)}} + \mathbf{m}_{\nu}\bigl(s^{(k)} \bigr) \bigr\|^2 ds \biggr) \biggr\} \, d\gamma_{\nu} \, P, \end{aligned}$$
$$\begin{aligned} \varGamma^{k}_{\nu,t}(\mu) =& \int_{\mathcal{C}} \log\biggl( \int\exp\biggl\{ \int_0^t \bigl( \mathbf{G}_{s^{(k)}}(\omega)+\mathbf{m}_{\nu }\bigl(s^{(k)} \bigr) \bigr)' \cdot d\mathbf{W}_s(x) \\ &{}- \frac{1}{2} \int_0^t \bigl\| \mathbf{G}_{s^{(k)}}(\omega)+\mathbf{m}_{\nu } \bigl(s^{(k)}\bigr) \bigr\| ^2ds \biggr\} d\gamma_{K^t_{\nu}}( \omega) \biggr) d\mu(x), \end{aligned}$$

as well as \(Q_{\nu,t}^{\alpha,k}\) and \(\varGamma^{\alpha, k}_{\nu,t}\) where

$$\begin{aligned} K_{\mu}^t(s,u)= \Biggl( \mathbf{1}_{\alpha=\beta} \displaystyle{\frac {1}{\lambda_{\alpha}^2} \sum_{\gamma=1}^M \sigma_{\alpha\gamma }^2 \int_{\mathcal{C}} S_{\alpha\gamma}\bigl(x^{\gamma}_{s-\tau_{\alpha\gamma} }\bigr)S_{\alpha\gamma} \bigl(x^{\gamma}_{u-\tau_{\alpha\gamma}}\bigr) d\mu(x)} \Biggr)_{\alpha , \beta\in\{1 \cdots M\}}. \end{aligned}$$

is define on [0,t]2.

These functions are clearly continuous in time on [0,T]. The result \(I(Q_{\nu,\tau_{m}\wedge T}^{k}|P) < \infty\) obviously remains.

By Jensen inequality, we have

$$\begin{aligned} \frac {\mbox {d} Q_{\nu,\tau^{\alpha}_{m}\wedge T}^{\alpha, k}}{\mbox {d} P_{\alpha }} (x) \geq&\exp\biggl \{-\frac{k_{\alpha}(\tau^{\alpha}_{m}(x)\wedge T)}{2\lambda _{\alpha}^2} \biggr\} \exp\biggl\{-\frac{\bar{J}_{\alpha}^2 (\tau ^{\alpha}_{m}(x)\wedge T)}{2 \lambda_{\alpha}^2} \biggr\} \\ &{}\times \exp \biggl\{\int_0^{\tau^{\alpha}_{m}(x)\wedge T} m^{\alpha}_{\nu} \bigl(t^{(k)}\bigr) dW^{\alpha }_t(x) \biggr\} \\ \geq&\exp\biggl\{-\frac{(k_{\alpha}+ \bar{J}_{\alpha}^2) T}{2\lambda_{\alpha}^2}-m \biggr\}, \end{aligned}$$

so that

$$\begin{aligned} \frac {\mbox {d} Q_{\nu,\tau_{m}\wedge T}^k}{\mbox {d} P} (x) \geq\exp\Biggl\{-\sum _{\alpha =1}^M \frac{(k_{\alpha}+ \bar{J}_{\alpha}^2) T}{2\lambda_{\alpha }^2}-Mm \Biggr\}. \end{aligned}$$

We then apply the same proof as in [3, Appendix B] to find:

$$\begin{aligned} \forall\mu\in\mathcal{M}_1^+(\mathcal{C}), \ H_{\nu,\tau _{m}\wedge T}^k = I\bigl(\mu|Q_{ \nu,\tau_{m}\wedge T}^k \bigr). \end{aligned}$$

Letting m to infinity, we conclude using the continuity of \(H_{\nu ,t}^{k}\) and \(I(.|Q_{\nu,t}^{k})\) on [0,T].

Proof of Lemma 5.(vi)

In order to demonstrate that H k is a good rate function, we need to show that it is lower semi-continuous and that it has compact level sets, i.e. {H kL} is a compact set for any L>0. This is a direct consequence of points (i)–(iv) proved above. □

Appendix B: Large Deviation Principle: Proof of the Technical Lemma 2

This appendix is concerned with the proof of Lemma 2 ensuring an exponential bound that will be used to show a tightness result on the sequence of empirical laws.

Proof

Let

$$\begin{aligned} B^n = \int_{\hat{\mu}_n \in B(\nu,\delta)} \exp\bigl\{ a n \bigl( \varGamma(\hat{\mu}_n)- \varGamma_{\nu}(\hat{ \mu}_n) \bigr) \bigr\} dQ_{\nu}^n. \end{aligned}$$

We introduce the probability measure Q ν defined on \(\mathcal {C}\) by:

$$\begin{aligned} \frac{dQ_{\nu}}{dP}(x) = \exp\bigl(\varGamma_{\nu}( \delta_x)\bigr) =& \int\exp \biggl(\int_0^T \bigl(\mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr)' \cdot d\mathbf{W}_t(x) \\ &{} - \frac {1}{2}\int _0^T \bigl\|\mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr\|^2 dt \biggr) d\gamma_{\nu }, \end{aligned}$$

so that \(Q_{\nu}^{n}=Q_{\nu}^{\otimes n}\). Writing the definitions of Γ and Γ ν , we find:

$$\begin{aligned} B^n = \int_{\hat{\mu}_n \in B(\nu,\delta)} \prod _{i=1}^n \biggl( \frac {{\mathcal{E}}_{\hat{\mu}_n} [\exp\{ \int_0^T ( \mathbf{G}_t + \mathbf{m}_{\hat {\mu}_n}(t) )' \cdot d\mathbf{W}^i_t -\frac{1}{2} \int_0^T \| \mathbf{G} _t + \mathbf{m} _{\hat{\mu}_n}(t) \|^2 dt \} ]}{{\mathcal{E}}_{\nu } [\exp \{ \int_0^T ( \mathbf{G}_t + \mathbf{m}_{\nu}(t) )' \cdot d\mathbf{W} ^i_t - \frac {1}{2} \int_0^T \| \mathbf{G}_t + \mathbf{m}_{\nu}(t) \| ^2 dt \} ]} \biggr)^a d(Q_{\nu})^{\otimes n}. \end{aligned}$$

Let ξ be a probability measure on \(\mathcal{C}\times\mathcal {C}\) with marginals \(\hat{\mu}_{n}\) and ν. We then have:

$$\begin{aligned} B^n = \int_{\hat{\mu}_n \in B(\nu,\delta)} \prod _{i=1}^n \biggl( \frac {{\mathcal{E}}_{\xi} [\exp\{ \int_0^T ( \mathbf {G}_t + \mathbf {m}_{\hat{\mu }_n}(t) )' \cdot d\mathbf{W}^i_t - \frac{1}{2} \int_0^T \| \mathbf{G}_t + \mathbf{m} _{\hat{\mu}_n}(t) \|^2 dt \} ]}{{\mathcal{E}}_{\xi } [\exp \{ \int_0^T ( \mathbf{G}_t' + \mathbf{m}_{\nu}(t) )d\mathbf{W}^i_t - \frac{1}{2} \int_0^T \| \mathbf{G}_t' + \mathbf{m}_{\nu}(t) \|^2 dt \} ]} \biggr)^a d(Q_{\nu})^{\otimes n} \end{aligned}$$

where (G,G′) is a 2M-dimensional Gaussian centered process with covariance K ξ (see (34)).

$$\begin{aligned} K_{\xi}(s,t) = \left ( \begin{array}{c@{\quad}c} K_{\mu}(s,t) & (\mathbf{1}_{\{\alpha=\gamma\}} K^{\alpha }_{\xi }(s,t) )_{\alpha, \gamma= 1 \cdots M} \\ (\mathbf{1}_{\{\alpha=\gamma\}} K^{\alpha}_{\xi}(s,t) )_{\alpha, \gamma= 1 \cdots M} & K_{\nu}(s,t) \\ \end{array} \right ) . \end{aligned}$$
(44)

Let

Then

$$\begin{aligned} B^n = &\int_{\hat{\mu}_n \in B(\nu,\delta)} \prod _{i=1}^n \biggl( \frac {{\mathcal{E}}_{\xi} [\exp\{ Y_i \} ]}{{\mathcal {E}}_{\xi} [\exp \{ Y_i' \} ]} \biggr)^a d(Q_{\nu})^{\otimes n} \\ = &\int_{\hat{\mu}_n \in B(\nu,\delta)} \prod_{i=1}^n \biggl( {\mathcal{E}}_{\xi } \biggl[ \frac{\exp{Y_i'}}{{\mathcal{E}}_{\xi} [\exp{Y_i'} ]} \exp{ \bigl( Y_i - Y_i' \bigr) } \biggr] \biggr)^a d(Q_{\nu})^{\otimes n} \\ \leq&\int_{\hat{\mu}_n \in B(\nu,\delta)} \prod_{i=1}^n {\mathcal{E} }_{\xi} \biggl[ \frac{\exp{Y_i'}}{{\mathcal{E}}_{\xi} [\exp { Y_i'} ]} \exp{ a \bigl( Y_i - Y_i' \bigr) } \biggr] d(Q_{\nu})^{\otimes n} \end{aligned}$$

by Jensen inequality.

Then, using Holder inequality twice with conjugate exponents (p,q) and (σ,η), one finds:

$$\begin{aligned} B^n \leq&\Biggl\{ \overbrace{\int\prod_{i=1}^n \frac{{\mathcal {E}}_{\xi} [\exp{ pY_i'} ]}{ ( {\mathcal{E}}_{\xi} [\exp{Y_i'} ] )^p } d(Q_{\nu})^{\otimes n}}^{B^n_1} \Biggr \}^{\frac{1}{p}} \biggl\{ \overbrace{\int\exp\bigl\{n \sigma \varGamma_{\nu}(\hat{\mu}_n) \bigr\} dP^{\otimes n}}^{B^n_2} \biggr\}^{\frac{1}{q\sigma}} \\ &{}\times\Biggl\{ \underbrace{\int_{\hat{\mu}_n \in B(\nu,\delta)} \prod _{i=1}^n {\mathcal{E}}_{\xi} \bigl[ \exp{ a \eta q \bigl( Y_i -Y_i' \bigr) } \bigr] dP^{\otimes n}}_{B^n_3} \Biggr\}^{\frac{1}{q\eta}}. \end{aligned}$$
(45)

We first bound the first term of the right hand side. Let \(\gamma _{p,\widetilde{K}^{T}_{\mu}}\) be a probability measure on Ω such that \(d\gamma_{p,\widetilde{K}^{T}_{\mu}} = \frac{\prod_{\gamma =1}^{M} \exp {-\frac{p}{2}\int_{0}^{T} {G^{\gamma}_{t}}^{2} dt} }{\int\prod_{\gamma=1}^{M} \exp{-\frac{p}{2}\int_{0}^{T} {G^{\gamma}_{t}}^{2} dt} d\gamma_{\mu} } d\gamma _{\mu}\). According to appendix A of [3] (where p=β 2), we have, for any p≥0, that G is a M-dimensional centered Gaussian process under \(\gamma_{p,\widetilde {K}^{T}_{\mu}}\).

Consequently, using the independence of (G,G′)’s components:

$$\begin{aligned} &{\mathcal{E}}_{\xi} \bigl[\exp{ pY_i'} \bigr] \\ &\quad{} = \prod_{\alpha =1}^M \exp\biggl\{ p \biggl( \int_0^T m_{\nu}^{\alpha}(t) dW^{i_{\alpha}}_t - \frac {1}{2} \int_0^T {m^{\alpha}_{\nu}}^2(t) dt \biggr) \biggr\} \; { \mathcal{E}}_{\xi} \biggl[ \exp\biggl\{-\frac{p}{2} \int _0^T {{G^{\alpha}_t}'}^2 dt \biggr\} \biggr] \\ &\qquad{}\times\int\exp\biggl\{ p \biggl(\int_0^T {G^{\alpha}_t}' \bigl(dW^{i_{\alpha}}_t - m_{\nu}^{\alpha}(t) dt \bigr) \biggr) \biggr\} d\gamma _{p,\widetilde{K}^T_{\nu}} \\ &\quad{} = \prod_{\alpha=1}^M \exp\biggl\{ p \biggl( \int_0^T m_{\nu }^{\alpha}(t) dW^{i_{\alpha}}_t - \frac{1}{2} \int_0^T {m^{\alpha}_{\nu}}^2(t) dt \biggr) \biggr\} \; { \mathcal{E}}_{\xi} \biggl[ \exp\biggl\{-\frac{p}{2} \int _0^T {{G^{\alpha}_t}'}^2 dt \biggr\} \biggr] \\ &\qquad{}\times \exp\biggl\{ \biggl( \frac{p^2}{2} \int\biggl(\int _0^T G^{\alpha }_t \bigl(dW^{i_{\alpha}}_t - m_{\nu}^{\alpha}(t) dt \bigr) \biggr)^2 \frac {\exp { (-\frac{p}{2}\int_0^T {G^{\alpha}_t}^2 dt )} }{\int \exp { (-\frac{p}{2}\int_0^T {G^{\alpha}_t}^2 dt )} d\gamma _{\nu}} d\gamma_{\nu} \biggr) \biggr\} . \end{aligned}$$

Hence,

$$\begin{aligned} \frac{{\mathcal{E}}_{\xi} [ \exp{pY_i'} ]}{{\mathcal {E}}_{\xi} [ \exp {Y_i'} ]^p} = &\prod_{\alpha=1}^M \underbrace{\frac{{\mathcal{E}}_{\xi } [ \exp {-\frac{p}{2} \int_0^T {{G^{\alpha}_t}'}^2 dt} ]}{{\mathcal {E}}_{\xi } [ \exp{-\frac{1}{2} \int_0^T {{G^{\alpha}_t}'}^2 dt} ]^p}}_{f^{\alpha }(p)} \exp\biggl\{ \frac{p}{2} \int\biggl(\int_0^T G^{\alpha}_t \bigl(dW^{i_{\alpha}}_t - m_{\nu}^{\alpha}(t) dt \bigr) \biggr)^2 \\ &{} \times\underbrace{ \biggl(p\frac{\exp{-\frac{p}{2}\int_0^T {G^{\alpha }_t}^2dt} }{\int\exp{-\frac{p}{2}\int_0^T {G^{\alpha}_t}^2dt} d\gamma _{\nu}} - \varLambda_T \bigl(G^{\alpha}\bigr) \biggr)}_{g^{\alpha}(p)} d\gamma_{\nu} \biggr\}. \end{aligned}$$

Remark thatf α(p) is bounded for p>0. In fact, on one hand it is clear that

$$\begin{aligned} \forall p \in[0,+\infty[, {\mathcal{E}}_{\xi} \biggl[ \exp{- \frac {p}{2} \int_0^T {{G^{\alpha}_t}'}^2 dt} \biggr]\leq1, \end{aligned}$$

furthermore Jensen inequality and Fubini theorem give us:

$$\begin{aligned} {\mathcal{E}}_{\nu} \biggl[\exp\biggl\{-\frac{p}{2}\int _0^T {G^{\alpha}_t}^2 dt \biggr\} \biggr] \geq\exp\biggl\{ - \frac{p}{2} \int _0^T {\mathcal{E}}_{\nu } \bigl[ {G^{\alpha }_t}^2 \bigr] dt \biggr\} \geq\exp\biggl \{ -\frac{p T k_{\alpha}}{2 \lambda_{\alpha}^2} \biggr\} . \end{aligned}$$

Therefore, bounded convergence monotone gives f α(p)→1 and, similarly, g α(p)→0 as p↘1. Moreover, as \((\int_{0}^{T} {G^{\alpha}_{t}}' (dW^{i_{\alpha}}_{t} - m_{\nu}^{\alpha}(t) dt ) )^{2}\) has finite moments under γ ν , we can find a finite constant C 1(p), C 1(p)↘0 as p↘1, such that:

$$\begin{aligned} B^n_1 = \biggl( \int\frac{{\mathcal{E}}_{\xi} [\exp{ pY_i'} ]}{ ( {\mathcal{E} }_{\xi} [\exp{ Y_i'} ] )^p } dQ_{\nu} \biggr)^n \leq e^{ C_1(p) n} . \end{aligned}$$
(46)

Moreover:

$$\begin{aligned} B^n_2 = & \biggl( \int{\mathcal{E}}_{\nu} \biggl[ \exp\biggl\{ \int_0^T \bigl( \mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr)' \cdot d\mathbf{W}_t(x) - \frac{1}{2} \int_0^T \bigl\|\mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr\|^2 dt \biggr\} \biggr]^{\sigma} dP(x) \biggr)^n \\ \leq&\biggl( {\mathcal{E}}_{\nu} \biggl[ \int\exp\biggl\{ \sigma \int_0^T \bigl(\mathbf{G} _t + \mathbf{m}_{\nu}(t) \bigr)' \cdot d\mathbf{W}_t(x) \biggr\} dP(x)\\ &{}\times \exp\biggl\{ - \frac{\sigma}{2} \int_0^T \bigl\|\mathbf{G}_t + \mathbf{m}_{\nu }(t) \bigr\|^2 dt \biggr\} \biggr] \biggr)^n \\ = & \biggl( {\mathcal{E}}_{\nu} \biggl[ \exp\biggl\{ \frac{\sigma ^2-\sigma }{2} \int_0^T \bigl\| \mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr\|^2 dt \biggr\} \biggr] \biggr)^n. \end{aligned}$$

So that if we take σ close enough to 1, we can find a finite constant C 2(σ), lim σ→1 C 2(σ)=0, such that (see inequality (38)):

$$\begin{aligned} B^n_2 \leq\biggl( {\mathcal{E}}_{\nu} \biggl[ \exp\biggl\{ \frac {\sigma ^2-\sigma }{2} \int_0^T \bigl\| \mathbf{G}_t + \mathbf{m}_{\nu}(t) \bigr\|^2 dt \biggr\} \biggr] \biggr)^n < e^{ C_2(\sigma) n } . \end{aligned}$$
(47)

We now will bound the last term of the right hand side of (45). By Cauchy-Schwarz inequality, if κ=qaη:

$$\begin{aligned} B^n_3 \leq&\Biggl\{ \int_{\hat{\mu}_n \in B(\nu,\delta)} \prod_{i=1}^n {\mathcal{E}}_{\xi} \biggl[ \exp\biggl\{ 2\kappa\int_0^T \bigl( \mathbf{G}_t-\mathbf{G} _t'+ (\mathbf{m} _{\hat{\mu}_n}-\mathbf{m}_{\nu}) (t) \bigr)' \cdot d \mathbf{W}^i_t \\ &{}- 2\kappa^2 \int_0^T \bigl \| \mathbf{G}_t - \mathbf{G}_t' + (\mathbf{m}_{\hat {\mu}_n} -\mathbf{m}_{\nu}) (t) \bigr\|^2 dt \biggr\} \biggr] dP^{\otimes n} \Biggr\} ^{\frac{1}{2}} \\ &{}\times \biggl\{ \int_{\hat{\mu}_n \in B(\nu,\delta )} {\mathcal{E} }_{\xi} \biggl[ \exp \biggl\{ 2\kappa^2 \int_0^T \bigl\| \mathbf{G}_t - \mathbf{G}_t' + ( \mathbf{m}_{\hat{\mu}_n} - \mathbf{m}_{\nu }) (t) \bigr\|^2 dt \\ &{} - \kappa\int_0^T \bigl\|\mathbf{G}_t+ \mathbf{m}_{\hat{\mu }_n}(t) \bigr\| ^2 - \bigl\| \mathbf{G} _t' + \mathbf{m}_{\nu}(t) \bigr\|^2 dt \biggr\} \biggr]^n dP^{\otimes n} \biggr\} ^{\frac{1}{2}}. \end{aligned}$$

The first term is the square root of a martingale’s expectation, thus equal to one. For the second term, we remark that:

$$\begin{aligned} &- \int_0^T \bigl(G^{\alpha}_t+m^{\alpha}_{\hat{\mu}_N}(t) \bigr)^2 - \bigl({G^{\alpha}_t}' + m_{\nu}^{\alpha}(t) \bigr)^2 dt\\ &\quad{} \leq \frac{\delta ^{\frac{1}{2}}}{2} \biggl( \frac{1}{\delta} \int_0^T \bigl(G^{\alpha}_t- {G^{\alpha}_t}' + \bigl(m^{\alpha}_{\hat{\mu}_N}(t) - m_{\nu}^{\alpha } \bigr) (t) \bigr)^2 dt \\ &\qquad{}+ \int_0^T \bigl(G^{\alpha}_t+ {G^{\alpha}_t}' + \bigl(m^{\alpha}_{\hat {\mu }_N}(t)+m^{\alpha}_{\nu} \bigr) (t) \bigr)^2 dt \biggr), \end{aligned}$$

so that, by Cauchy-Schwarz inequality,

$$\begin{aligned} B^n_3 \leq&\biggl\{ \int{\mathcal{E}}_{\xi} \biggl[ \exp\biggl\{ \bigl(4\kappa^2+\kappa\delta^{-\frac{1}{2}} \bigr) \int_0^T \bigl\| \mathbf{G}_t- \mathbf{G} _t'+ (\mathbf{m} _{\hat{\mu}_n}- \mathbf{m}_{\nu}) \bigl(t^{(k)}\bigr) \bigr\|^2 dt \biggr\} \biggr]^n dP^{\otimes n} \biggr\}^{\frac{1}{4}} \\ &{} \times\biggl\{ \int{\mathcal{E}}_{\xi} \biggl[ \exp\biggl\{ \kappa \delta^{\frac {1}{2}}\int_0^T \bigl\| \mathbf{G}_t+\mathbf{G}_t'+ (\mathbf {m}_{\hat{\mu }_n}+\mathbf{m}_{\nu}) (t) \bigr\|^2 dt \biggr\} \biggr]^n dP^{\otimes n} \biggr\}^{\frac{1}{4}} \\ \leq&\exp{\sum_{\alpha=1}^n \biggl\{ \frac{1}{2} \bigl(4\kappa^2+\kappa\delta^{-\frac{1}{2}} \bigr)\frac{\bar{J}_{\alpha}^2 K_S^2}{\lambda_{\alpha }^2}T\delta^2 + \biggl(2\kappa\delta ^{\frac{1}{2}}T\frac{\bar{J}_{\alpha}^2}{\lambda_{\alpha}^2}\biggr) \biggr\}} \\ &{} \times\biggl\{ \int{\mathcal{E}}_{\xi} \biggl[ \exp\biggl\{ 2 \bigl(4\kappa^2+\kappa\delta^{-\frac{1}{2}} \bigr) \int _0^T \|\mathbf{G}_t-\mathbf{G} _t' \|^2 dt \biggr\} \biggr]^n dP^{\otimes n} \biggr\}^{\frac{1}{4}} \\ & {}\times\biggl\{ \int{\mathcal{E}}_{\xi} \biggl[ \exp\biggl\{ 2 \kappa\delta^{\frac {1}{2}}\int_0^T \| \mathbf{G}_t+\mathbf{G}_t' \|^2 dt \biggr\} \biggr]^n dP^{\otimes n} \biggr\}^{\frac{1}{4}}. \end{aligned}$$

As \({\mathcal{E}}_{\xi} [ \int_{0}^{T} \|\mathbf{G}_{t}-\mathbf {G}_{t}' \|^{2} dt ] = \sum_{\alpha, \gamma=1}^{M} \frac{\sigma_{\alpha\gamma}^{2} K_{S}^{2}}{\lambda_{\alpha} ^{2}} \int_{0}^{T} \int| S_{\alpha\gamma}(x^{\gamma}_{t-\tau_{\alpha\gamma}}) - S_{\alpha \gamma}(y^{\gamma}_{t-\tau_{\alpha\gamma}}) |^{2} d\xi (x,y) dt \leq\sum_{\alpha, \gamma=1}^{M} \frac{\sigma_{\alpha \gamma}^{2} T}{\lambda_{\alpha}^{2}} \delta ^{2}\) (in fact, \(d_{T}(\hat{\mu}_{n},\nu) \leq\delta\)), we can use Appendix A, Lemma 3.2 of [3] to bound the two last term of the previous inequality: there exists two finite constants \(C^{\kappa }_{1}(\delta)\) and \(C^{\kappa}_{2}(\delta)\) such that

Hence, we can find C κ (δ),lim δ→0 C κ (δ)=0, such that

$$\begin{aligned} B^n_3 \leq e^{C_{\kappa}(\delta) n} . \end{aligned}$$
(48)

We conclude by using (46), (47) and (48) in (45). □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Cabana, T., Touboul, J. Large Deviations, Dynamics and Phase Transitions in Large Stochastic and Disordered Neural Networks. J Stat Phys 153, 211–269 (2013). https://doi.org/10.1007/s10955-013-0818-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-013-0818-5

Keywords

Navigation